You are on page 1of 18

Chem Soc Rev

View Article Online


TUTORIAL REVIEW View Journal | View Issue

Fabrication of desalination membranes by


This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

interfacial polymerization: history, current efforts,


Cite this: Chem. Soc. Rev., 2021,
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

50, 6290 and future directions


ab
Xinglin Lu and Menachem Elimelech *b

Membrane desalination is a promising technology for addressing the global challenge of water scarcity
by augmenting fresh water supply. Continuous progress in this technology relies on development of
membrane materials. The state-of-the-art membranes used in a wide range of desalination applications
are polyamide thin-film composite (TFC) membranes which are formed by interfacial polymerization (IP).
Despite the wide use of such membranes in desalination, their real-world application is still hampered
by several technical obstacles. These challenges of the TFC membranes largely stem from the inherent
limitations of the polyamide chemistry, as well as the IP reaction mechanisms. In the past decade, we
have witnessed substantial progress in the understanding of polyamide formation mechanisms and the
development of new IP strategies that can potentially lead to the redesign of TFC membranes. In this
Tutorial, we first present a brief history of the development of desalination membranes and highlight the
major challenges of the existing TFC membranes. We then proceed to discuss the pros and cons of
emerging IP-based fabrication strategies aiming at improving the performance of TFC membranes. Next,
Received 19th January 2021 we present technical obstacles and recent efforts in the characterization of TFC membranes to enable
DOI: 10.1039/d0cs00502a fundamental understanding of relevant mechanisms. We conclude with a discussion of the current gap
between industrial needs and academic research in designing high-performance TFC membranes, and
rsc.li/chem-soc-rev provide an outlook on future research directions for advancing IP-based fabrication processes.

Key learning points


(1) Current status of desalination processes and historical timeline for the development of desalination membranes.
(2) Major challenges of thin-film composite (TFC) membranes made by interfacial polymerization (IP) in desalination applications.
(3) Emerging strategies of IP for fabricating high-performance TFC membranes.
(4) Emerging techniques for polyamide thin-film characterization for revealing structural information of TFC membranes.
(5) Gap between industrial needs and academic research in designing high-performance TFC membranes, and potential research directions for advancing
IP-based fabrication processes.

1. Introduction is predicted to double in the coming decade. To address this global


challenge, we need technological innovations to enable an efficient
Water is a precious resource, essential to humanity’s sustainability and sustainable supply of fresh water.
and well-being.1 However, population growth, industrialization, Desalination can offer a steady supply of high-quality fresh
and water contamination have led to the pressing challenge of water through the treatment of seawater and brackish ground-
water scarcity around the globe.2 Presently, over one billion people water. These saline waters account for 97.5% of the water sources
do not have adequate access to safe drinking water, and this figure on our planet, making desalination an important solution to
addressing water scarcity on a global scale.3 To date, over 18 000
operational desalination plants provide B90 million m3 of fresh
a
CAS Key Laboratory of Urban Pollutant Conversion, Department of Environmental
water daily to meet the demand of 300 million people, a number
Science and Engineering, University of Science and Technology of China,
Hefei 230026, China
equivalent to the population of the United States.4
b
Department of Chemical and Environmental Engineering, Yale University, Growing interest in desalination has resulted in several
New Haven, Connecticut 06520-8286, USA. E-mail: menachem.elimelech@yale.edu technologies, which can be largely classified into two

6290 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev


This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

Fig. 1 Overview of the development and current status of desalination processes. (A) Trends in global desalination capacity in the past decades. (B) Share
of operational facilities using different technologies, including reverse osmosis (RO), nanofiltration (NF), multi-stage flash (MSF), and multi-effect
distillation (MED). Figures are reproduced with permission from ref. 6, Elsevier.

categories: thermal-based processes and filtration-based energy than thermal-based processes in seawater desalination,
processes. Thermal-based processes exploit a thermal-driven despite measures to recover the latent heat of vaporization in
phase change to separate water from non-volatile contaminants thermal desalination technologies. Consequently, filtration-
like salt. These processes, such as multi-effect distillation based RO has become the predominant destination technology,
(MED) and multi-stage flash (MSF), were widely employed in as illustrated by the rapid growth in the past two decades
the Gulf countries before the 1990s (Fig. 1A). Filtration-based (Fig. 1A), accounting for B70% of the global desalination
processes, including reverse osmosis (RO) or nanofiltration capacity (Fig. 1B).6
(NF), rely on external pressure to drive water molecules across The development of the RO process has largely been governed
a semi-permeable membrane. Although both thermal-based by advances in membrane materials, the core of the whole
and filtration-based processes are effective in producing fresh process. A chronological summary of the evolution of the RO
water, the intrinsic difference in driving force leads to different desalination process is presented in Fig. 2 to illustrate the major
separation principles and energy-consumption levels, which milestones in the development of this technology. While the
were explained in a recent review article.5 In brief, thermal phenomenon of osmosis was first discovered in 1748, the proof-
desalination processes must overcome the enthalpy of vapor- of-concept of RO desalination was not demonstrated until two
ization (i.e., latent heat, DHvap), while the filtration processes centuries later, in the 1950s, using symmetric cellulose acetate
must overcome the Gibbs free energy of separation (DGmix). To films.7 Despite good rejection of NaCl, the water flux of such
produce a unit volume of fresh water under typical operating films was too low to be practically viable. Later, in the 1960s, two
conditions, DHvap is at least two orders of magnitude larger pioneers—Loeb and Sourirajan—developed the first asymmetric
than DGmix (667 vs. 1.06 kW h1 m3 for 50% water recovery). cellulose acetate (CA) membrane via phase separation.7 This
As a result, filtration-based processes consume much less membrane exhibited high salt rejection with ten times larger

Xinglin Lu is a tenure-track Menachem Elimelech is the


assistant professor in the Sterling Professor of Chemical
Department of Environmental and Environmental Engineering
Science and Engineering at the at Yale University. His research
University of Science and interests include emerging
Technology of China (USTC). He membrane-based technologies at
received his PhD from Harbin the water-energy nexus, materials
Institute of Technology under for next generation desalination
the supervision of Prof. Jun Ma, and water purification
followed by postdoctoral training membranes, and environmental
in Prof. Menachem Elimelech’s applications of nanomaterials.
group at Yale University. He is Professor Elimelech is a
Xinglin Lu broadly interested in Menachem Elimelech Clarivate Analytics (formerly
technologies that hold potential Thomson Reuters) Highly Cited
in addressing challenges at the water-energy nexus. His current Researcher in two categories: chemistry and environment/ecology.
research focuses on advanced membrane materials and processes He is a member of the US National Academy of Engineering and a
for sustainable water supply and wastewater management. foreign member of the Chinese Academy of Engineering.

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6291
View Article Online

Chem Soc Rev Tutorial Review


This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

Fig. 2 Historical timeline of the development of membrane-based desalination processes.


Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

water flux than that of any available membranes, thereby facilitating Although the advent of TFC membranes resulted in remarkable
real-world applications of the RO process. Using this asymmetric energy savings in desalination, certain features of the fabrication
cellulose acetate membrane, the first pilot-scale RO system procedure and film properties are inherently limiting. For
was demonstrated in California in the 1960s, followed by the example, the seemingly straightforward IP process turns out to
establishment of the first RO municipal plants in Florida in the be rather complex, limiting the fundamental understanding of
1970s.7 These practical applications also enabled engineers and polyamide formation mechanisms. As such, existing industrial
researchers to realize the shortcomings of the cellulose-type fabrication processes are largely empirical and lack molecular-level
membranes in long-term operation, including a narrow operating design, leading to the formation of TFC membranes with
pH range and susceptibility to biological degradation. To uncontrolled characteristics. Some membrane characteristics lead
overcome these limitations, Cadotte developed the first thin-film to deleterious phenomena (e.g., fouling and low rejection of
composite (TFC) polyamide membranes in the late 1970s.8 neutral solutes) that are detrimental to product water quality
Compared with the cellulose acetate membranes, the TFC and system efficiency. Given the dominant role of TFC membranes
membrane is characterized by exceptional desalination perfor- in RO desalination, there is a crucial need for a better mechanistic
mance (high water flux and high salt rejection) and enhanced understanding of the IP process to enable innovative designs and
stability under a wider pH range, thereby becoming the gold to further advance RO desalination technologies.
standard in RO-based desalination technologies.3 Herein, we review the most recent progress in the IP process
TFC membranes are fabricated through interfacial polymer- towards the fabrication of high-performance TFC membranes.
ization (IP)—a process that allows for the facile synthesis We begin by highlighting the major limitations of TFC membranes
of an ultrathin selective layer. A representative procedure for regarding their fabrication process and real-world applications. We
fabricating TFC RO membranes is illustrated in Fig. 3. proceed to analyze the pros and cons of emerging IP-based
A microporous polysulfone membrane is first immersed in an fabrication strategies that aim at improving the robustness of
aqueous diamine solution, m-phenylenediamine (MPD), to TFC membranes. We then present technical obstacles and recent
allow diamine penetration into the support. After removing efforts in the characterization of TFC membranes to enable
the excess diamine on the surface, the impregnated support is fundamental understanding of relevant mechanisms. Finally, we
brought into contact with an organic phase of acyl chloride— discuss the current gap between industrial needs and academic
trimesoyl chloride (TMC). As MPD monomers diffuse to the inter- research in fabricating TFC membranes, and provide an outlook
face between water and organic phases, they react with the TMC to on future research directions for advancing IP-based fabrication
form a thin polyamide layer on top of the polysulfone support. processes.

Fig. 3 Schematic representation of a typical interfacial polymerization process for making TFC RO membranes. The reaction between m-phenylene-
diamine (MPD) and trimesoyl chloride (TMC) occurs at the surface of a microporous polysulfone support to form a thin polyamide layer, whose chemical
formula is illustrated. The m and n in the polymer structure represent the crosslinked and the linear segments, respectively (m + n = 1).

6292 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev

2. Current challenges in fabricating Js = BDC (2)


high-performance TFC membranes where B is the solute permeability coefficient and DC is the
2.1. Breaking the trade-off between permeability and concentration difference across the selective layer.
selectivity For a given membrane, the water and salt permeability
coefficients of the active layer are determined by the intrinsic
Mass transport in the dense, nonporous polyamide film of TFC
water permeability (Pw) and solute permeability (Ps), respectively:9
membranes is governed by the solution-diffusion model. Water
and solute molecules first partition into the polyamide layer, Pw Vw
A¼ (3)
diffuse through the polymer matrix under chemical potential dm Rg T
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

gradient, and desorb into the permeate side. Separation of


fresh water from saline water is achieved because of the Ps
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

B¼ (4)
difference in the permeabilities of water and solutes in the dm
polymer matrix. Water flux, Jw, is expressed as
where Vw is the molar volume of water, dm is the film thickness, Rg
Jw = A(DP  Dp) (1) is the ideal gas constant, and T is the absolute temperature.
When plotting water–salt selectivity (Pw/Ps) versus water
where A is the water permeability coefficient (or permeance), permeability (Pw) in Fig. 4A, performance data of existing TFC
and DP and Dp are the applied pressure and osmotic pressure membranes are mostly confined in the blue-box region,
differences across the active layer, respectively. Similarly, solute wherein the highest selectivity for a given permeability
flux, Js, is determined by approaches the solid line. This line represents the trade-off

Fig. 4 Major challenges of interfacial polymerization and TFC membranes. (A) Permeability–selectivity trade-off relationship for TFC desalination
membranes. Performance of current TFC membranes mostly lays in the blue box, in which the highest water–salt selectivity at a given water permeability
is near the indicated trade-off line. Ideal membranes would have both high permeability and high selectivity, thereby breaking the trade-off relationship
and laying in the top-right corner of the plot (red box). However, this combination is difficult to attain. A more practical strategy is to fabricate TFC
membranes with comparable water permeability of current membranes but with increased water–salt selectivity (green box); this strategy can improve
the efficiency of current desalination processes. (B) Correlation of crosslinking degree with salt rejection and permselectivity of TFC membranes.
(C) Surface characteristics that determine fouling propensity of TFC membranes. (D) Proposed mechanisms for the origin of the characteristic ridge-and-
valley structure of TFC membrane during interfacial polymerization: generation of interfacial instability in the organic phase and formation of nano-size
bubbles in the aqueous phase.

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6293
View Article Online

Chem Soc Rev Tutorial Review

relationship between permeability and selectivity, which is polyamide forms a fully crosslinked structure (i.e., the ‘‘m’’
widely observed in membrane processes using synthetic segment in Fig. 3) with a monomer ratio of 3 : 2 for MPD to
polymer materials.10 Ideal membranes should have high values TMC. In an actual IP process, some linear segments (i.e., the
of both water permeability and water–salt permselectivity ‘‘n’’ segment in Fig. 3) would also form under a monomer ratio
(i.e., the red-box region in Fig. 4A). Nevertheless, such a of 1 : 1 for MPD to TMC. In this case, unreacted acyl chloride
combination is difficult to achieve, as most material properties groups undergo hydrolysis to form carboxylic groups, thereby
that affect water permeability would in turn exert similar effects imparting a negative charge to the membrane surface. The
on solute permeability. Indeed, the influence of material crosslinked segments more rigidly resist swelling due to the
properties on salt permeability is more pronounced as salt presence of chain linkages between polyamide backbones, thus
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

permeability is proportional to the cube of water permeability impeding solute transport. In contrast, the linear segments are
(Ps p Pw3).9,11 That is, any increment in water permeability likely to undergo more severe swelling, which facilitates solute
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

would lead to an exponential increase in salt permeability and transport within the polymer matrix. As shown in Fig. 4B, both
thus a remarkable decrease in selectivity. Consequently, despite NaCl rejection and permselectivity exhibit an overall good
substantial efforts in optimization of IP reaction conditions, correlation with the crosslinking density, i.e., m/(m + n),
progress in pushing the trade-off line toward the ideal regime implying that increasing crosslinking segments is a feasible
(i.e., the red arrow in Fig. 4A) is relatively slow. strategy toward the design of membrane with high water–salt
Instead of attaining membranes with high values of both selectivity. We note that this is only one example of how material
water permeability and water–salt selectivity, recent studies characteristics determine membrane transport properties. Efforts
emphasized the critical need for increased water–salt selectivity to elucidate the structure–property–performance relationship of
rather than increased permeability for the design of TFC the polyamide selective layer will be of paramount importance in
membranes.11,12 Based on process modeling of the desalination guiding the future design of high-performance TFC membranes.
process, it was shown that further increments of water permeability
would have marginal influence on the energy efficiency of RO 2.2. Developing fouling-resistant membranes
systems. For example, in seawater desalination, a 5-fold increase of Membrane materials effectively reject solutes to yield high-
water permeance of existing TFC membranes would only lead to quality product water. During operation, the rejected solutes
B3.7% reduction in energy consumption.11 Increasing water–salt continuously accumulate on the membrane surface, leading to
selectivity is important because RO exhibits relatively poor membrane fouling. Fouling is widely considered the Achilles
rejection of small neutral molecules, such as boron and heel for membrane processes as it leads to reduced process
disinfection byproducts, thus necessitating post-treatment performance. For instance, the presence of a foulant layer
processes for their complete removal. These additional increases the resistance for mass transport, resulting in a
processes, in turn, lead to increased capital and operating continuous decrease in water flux. Periodic physical or
costs, which could be avoided by using membranes with higher chemical cleaning is thus needed to mitigate fouling, leading
water–salt selectivity. Thus, increasing the water–salt selectivity to frequent system shutdown and increased operating cost.
of existing TFC membranes is a more effective strategy (the Therefore, understanding fouling mechanisms and developing
green arrow in Fig. 4A) for improving system efficiency and effective antifouling strategies are of critical importance.
product water quality. Readers are referred to a recent comprehensive review to gain
Water–solute selectivity of TFC membranes is determined by more insights into membrane fouling.14
the combined mechanisms of steric hindrance, Donnan exclusion, In seawater desalination, membrane fouling can be classified
and dielectric effect.13 Steric hindrance is related to the pore size into three categories (Fig. 4C): fouling by organic substances
(or free volume) of polyamide, which allows molecules with (e.g., natural organic matter), scaling by inorganic species (e.g.,
smaller sizes to permeate while keeping larger ones from doing gypsum or silica), and biologic fouling by microorganisms (e.g.,
so. Donnan exclusion is governed by the fixed charge density on bacteria or algae). The extent of fouling is governed by both
the membrane surface. For example, a negatively charged surface membrane surface properties and feedwater solution chemistry. The
effectively rejects anions (e.g., Cl). To maintain electroneutrality surface properties of MPD-based TFC membranes—hydrophobic
on both sides of the membranes, cations (e.g., Na+) are therefore nature, roughness, and presence of carboxylic groups—are
also excluded by the membrane, leading to a good overall rejection conducive to membrane fouling. Accordingly, surface modifi-
of salts (e.g., NaCl). The dielectric effect is related to an energetic cations that render the membrane surface hydrophilic, smooth,
penalty that solutes have to pay while transferring from a solvent and of neutral charge, have been widely proposed as effective
with a high dielectric constant (i.e., water) to a medium with a low antifouling strategies.15,16 In addition to surface modification,
dielectric constant (i.e., polymer membrane material). The fouling control could be achieved by adjusting solution
presence of such an energetic penalty thus hampers the diffusion chemistry prior to membrane desalination. For example, a large
of solutes (e.g., salt ions) through the membrane. portion of organic and inorganic species could be removed in
The three solute rejection mechanisms discussed above are pretreatment processes by adding coagulants or adjusting pH,
relevant to the material characteristics of the selective layer, thereby reducing the overall fouling potential of the feedwater.3
which stems from the molecular structure of the polyamide Compared with other types of fouling, biofouling is even
materials. Taking crosslinking degree as an example, the more challenging, owing to the growth and proliferation of

6294 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev

bacteria to form a sticky biofilm. Although the use of generation of byproduct hydrochloric acid in the TMD–MPD
disinfectants (e.g., chlorine) could effectively inactivate micro- reaction, lead to the degassing of dissolved gas molecules (e.g.,
organisms, the chemical structure of polyamide is highly prone CO2, O2, and N2) in the aqueous phase (Fig. 4D, lower region).21
to degradation when exposed to oxidants like chlorine.17,18 These bubbles are then encapsulated by the nascent polyamide
Polyamide degradation by chlorine undergoes a two-step layer, leading to the formation of the ridge-and-valley structure.
reaction—N-chlorination and ring chlorination—resulting in However, only indirect evidence was obtained for supporting
irreversible damage to the polyamide selective layer and loss of these two mechanisms, thereby calling for more extensive
salt rejection properties.18 Therefore, complete removal of efforts to reveal the mechanisms of roughness formation.
chlorine in the feedwater is needed before membrane filtration.
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

Notably, after membrane filtration, the product water is often


re-chlorinated for potable supply. Such complicated de- and
3. Emerging strategies of interfacial
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

re-chlorination procedures decrease system efficiency remarkably, polymerization for fabricating TFC
thus necessitating the development of chlorine-resistant membranes
membranes for biofouling control.17,18 Among the proposed
strategies are the use of oxidation-resistant monomers, incor- Recent advances in thin-film fabrication processes and material
poration of sacrificial materials, or surface modification.18,19 synthesis techniques enable innovative re-design of the IP
Despite these efforts, to date, there is no commercial chlorine- process. These newly proposed IP strategies aim to endow TFC
resistant desalination membrane, highlighting the challenges membranes with desired physicochemical characteristics,
and the urgent need to develop such a membrane. thereby leading to performance gains in fouling resistance or
water–solute selectivity. Herein, we classify these emerging IP
2.3. Understanding the reaction mechanisms of interfacial strategies into four categories according to their underlying
polymerization design principles (Fig. 5) and discuss their respective advantages/
Membrane desalination is an interfacial separation process disadvantages for making high-performance TFC membranes.
where mass transport and fouling propensity are largely Additionally, we compare the performance of membranes made
governed by membrane surface characteristics. To date, most using these emerging strategies to state-of-the-art commercial
surface characteristics (e.g., wettability, charge, and functional TFC membranes. Last, we discuss the key limitations of these
groups) have been well investigated for their effects on IP strategies and general trends in the future development of IP.
membrane performance. Researchers have also developed
numerous strategies to tune these characteristics for fabricating 3.1. Constructing sacrificial interlayers on the support layer
high-performance membranes. In contrast, the origin of IP reaction is performed on a microporous support membrane
membrane surface roughness resulting from the IP process that is formed by the non-solvent-induced phase separation
and its role in membrane performance are relatively poorly (NIPS) process. In brief, a film of polymer dissolved in solvent is
understood. soaked in a non-solvent bath (e.g., water or ethanol). The
The surface morphology of TFC membranes has a characteristic exchange of solvent and non-solvent leads to the precipitation
ridge-and-valley structure, with polyamide protrusions of of the dissolved polymer into a solid thin film, thereby producing
B100 nm. Such rough structure may shield accumulated foulants a porous membrane structure.22 However, as NIPS is largely a
in the valley regions from being removed under the hydrodynamics stochastic process, precise control over the structure of the
of crossflow filtration, thereby leading to increased membrane microporous support is relatively challenging, necessitating more
fouling. In contrast, due to the presence of hollow cavities inside effective strategies for tuning characteristics of the support
the protrusions, the rough morphology also increases the prior to IP.
surface area for water transport, which is advantageous for TFC One strategy that attracts a growing interest is to construct a
membrane design. As such, attaining surface morphology that sacrificial interlayer on the microporous support (Fig. 5A). The
could simultaneously inhibit fouling (low roughness) and facilitate sacrificial interlayers are generally made from nanomaterials
water transport (high roughness) is challenging. (e.g., nanostrands, nanotubes, or metal–organic frameworks)
Precise control of surface morphology relies on understanding with unique physicochemical properties (e.g., shape, wettability,
the formation mechanisms of the rough features during IP. and charge), effectively modifying the support layer properties
Two mechanisms have been proposed for the formation of the before IP. For example, a layer of needle-like cadmium hydroxide
ridge-and-valley structure during the IP process: interfacial nanostrands was first deposited on a porous support.20 Using
instability and formation of nano-size bubbles. For interfacial the modified support for IP, the presence of this nanostrand
instability, previous studies have suggested that the intense layer moved the interfacial reaction away from the solid support,
exothermic reaction between TMC and MPD increases the local thereby resulting in a thin, smooth polyamide film.20 Although
temperature.20 Such temperature change at the interface this strategy was first demonstrated for membranes used in
results in convective flow in the organic phase which forms organic solvent separation,20 similar techniques have recently
the rough features (Fig. 4D, upper region). For the second been applied for membranes for seawater desalination.23 Different
mechanism—the formation of nanobubbles—studies suggest from the nanostrand layer, researchers deposited a composite
that the increased local temperature, together with the layer comprising a layer of carbon nanotubes (CNT) with a layer

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6295
View Article Online

Chem Soc Rev Tutorial Review


This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

Fig. 5 Schematic diagrams of different types of emerging interfacial polymerization (IP) strategies for fabricating high-performance TFC membranes. (A)
Use of sacrificial interlayers for tuning support layer properties prior to IP. A sacrificial layer is first deposited on a substrate (upper panel). IP is performed
on the sacrificial-layer tailored substrate to form a polyamide layer on top (middle panel). The sacrificial layer is then removed by dissolving in a solvent
(lower panel). (B) Embedding of nanomaterials into the polyamide layer for obtaining a mixed-matrix structure. Nanofillers include nanoparticles,
nanotubes, nanosheets, and aquaporin vesicles, representing nanomaterials with 0D to 3D structures. (C) Tuning reaction conditions of IP by adding
different additives. In the upper left panel, the addition of polyvinyl alcohol (PVA) to the aqueous phase reduces the diffusion rate of PIP, leading to a
diffusion-driven instability and generating nanoscale spotted and striped Turing structures. In surfactant assembly regulated interfacial polymerization
(SARIP) (upper right panel), the addition of a surfactant, sodium dodecyl sulfate (SDS), results in a self-assembled network of amphiphiles at the water/
hexane interface, which facilitates the transport of PIP (red dots) across the interface to react with TMC (black dots). (D) Incorporation of new surface
coating techniques into IP processes. In electrospray process (upper panel), MPD (in aqueous phase) and TMC (in organic phase) solutions are
respectively introduced to two separate nozzles. A substrate stage is grounded and connected to the two nozzles with a direct current power source.
When high voltage is applied, the charged liquids leave the needles and emit microdroplets to deposit onto the substrate, where the reaction between
MPD and TMC leads to the formation of a thin polyamide film. In the molecular layer-by-layer process (lower panel), a substrate is alternately dipped into
the two monomer solutions (e.g., MPD in water and TMC in hexane) and subsequently rinsed with proper solvents (water and hexane) after each dipping
step. Such a dip-coating process is repeated for several cycles, resulting in a thin, defect-free film comprising multilayers of polyamide.

of zeolitic imidazolate framework (ZIF-8) on top.23 After the IP In addition to the use of nanomaterial-tailored supports,
reaction, the embedded ZIF-8 could be dissolved by immersing another strategy is to perform IP at a free interface between an
the membrane in water, leading to the formation of a hollow aqueous phase and an organic phase.25,26 The liquid interface
polyamide layer with crumpled surface structure. Such crumpled enables more rapid heat dissipation than that in an aqueous
structure resulted in an enhanced water permeability without solution trapped in solid support, forming free-standing
compromising salt rejection properties. Readers are referred to a nanofilms with smooth surfaces due to inhibition of interfacial
comprehensive review for the recent progress in the use of instability. The formed film could be used directly or
sacrificial interlayers in IP.24 transferred to another porous support for performance

6296 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev

evaluation. Such a strategy provides a useful platform for polyamide layer with favorable properties such as high water
isolating polyamide layers for advanced characterization in permeance and biofouling resistance. These attempts have
order to unravel the structure–property–performance relationship introduced the concept of thin-film nanocomposite (TFN)
of the polyamide selective layer. However, as the polyamide layer membranes; readers are referred to a comprehensive review to
is relatively fragile with low mechanical strength, integrating such learn more about TFN membranes.27 In the past decade, other
free-standing films into membrane modules will be extremely emerging materials with desired characteristics have also been
challenging. Moreover, even with an underlying support, the considered as embedded blocks in the polyamide layer. For
interaction between the polyamide and the support will be much example, researchers prepared lipid vesicles laden with aquaporins,
weaker compared with that of conventional TFC membranes. which have unique water channels with high water permeability
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

Such low interaction would result in low resistance of the and salt selectivity.28,29 The presence of the aquaporin water
polyamide films to high-pressure or high-crossflow conditions, channels facilitated water transport through the polyamide layer
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

causing the films to lose their integrity during operation. while retaining ionic species, and readers are referred to a recent
Therefore, further advances in material and process designs are comprehensive review.30
needed for expanding the real-world application of these free- Despite the reported performance gain in water permeance
standing polyamide membranes. and fouling resistance through adding filler materials, real-
world applications of such composite membranes are still
3.2. Embedding nanofillers in the polyamide active layer hindered by several challenges, casting doubt about their
Embedding fillers in the polyamide selective layer is, potentially, potential applicability. The most critical factor is the inevitable
a direct strategy toward high-performance TFC membrane formation of defects at the interface between fillers and the
fabrication (Fig. 5B); this strategy has been a topic of interest polyamide matrix, which has been a long-standing issue for all
for more than two decades.27 Following a typical IP process, types of dense membranes with a mixed-matrix structure.31 The
pre-synthesized fillers are added to the organic phase (mixed formation of these defects is either ascribed to the low
with TMC) or aqueous phase (mixed with MPD), forming a mix- miscibility between the two materials, or the inhibition of
matrix structure in the resulting polyamide layer. Early efforts the IP reaction to form less crosslinked polyamide regions.
largely focused on the use of different nanofillers (e.g., nano- For example, a recent study has observed the presence of
particles, nanotubes, and nanosheets), which endowed the non-selective defect channels with a size of B2 nm between

Fig. 6 Characteristics and performance of TFC membranes fabricated through emerging interfacial polymerization (IP) strategies. (A) Effect of
embedded nanofillers on the mass transport of TFC membranes. Cross-section image of a polyamide layer with embedded silver nanoparticles (AgNP)
(left panel). The light-colored haloes around the silver nanoparticles correspond to lower mass density, suggesting the possible formation of
nanochannels with a size of 2–3 nm between a silver nanoparticle and the polyamide matrix (right panel). These channels potentially facilitate water
permeability while compromising water–solute selectivity. Scanning electron microscopy is adapted with permission from ref. 32, American Chemical
Society. (B) Schematic illustration of spatial distribution of water permeability sites with different polyamide morphologies, including spotted (left panel)
and stripped (right panel) Turing structures. Figure is reproduced with permission from ref. 39, American Association for the Advancement of Science. (C)
Schematic illustrations of a polyamide selective layer formed via conventional IP (left), which has a heterogeneous pore size distribution with the
presence of large pores, and surfactant-assembly regulated IP (SARIP) (right), which has a uniform pore size distribution. The uniform pore structure
endows the SARIP membranes with a precise solute–solute separation.

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6297
View Article Online

Chem Soc Rev Tutorial Review

embedded silver nanoparticles and polyamide matrix morphology of the polyamide layer could be manipulated to
(Fig. 6A),32 which are detrimental to membrane water–salt form spotted (Fig. 6B, left panel) and striped (Fig. 6B, right
selectivity. Although modifying fillers with hydrophobic panel) Turing structures, thereby affecting the water transport in
materials may enhance their miscibility in the polyamide the resulting membranes. In another study,40 an opposite strategy
matrix and thus inhibit the formation of the defect was proposed to enhance the diffusion of PIP by adding surfactant
channels,33 the high liquid entry pressure (LEP) in these additives to the aqueous phase of diamine, leading to a new IP
hydrophobic nanopores would provide additional resistance process—surfactant assembly regulated interfacial polymerization
to water transport.34 For instance, a hydrophobic mesoporous (SARIP) (Fig. 5C, upper right panel).40 It was suggested that
material (contact angle of 1401) with a pore diameter of 4 nm the trans-interface diffusion of PIP in the conventional IP is a
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

requires an LEP of 138 bar, which is beyond the maximum rate-limiting step, resulting in a heterogeneous polyamide
operating pressure in typical RO systems (i.e., B80 bar). network (Fig. 6C, left panel) with a broad distribution of pore
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

Notably, the use of the fillers in desalination membranes may sizes. Addition of sodium dodecyl sulfate (SDS) molecules to
also result in the release of some filler materials to the product the aqueous phase in SARIP leads to the formation of a
water, leading to human exposure to these materials and self-assembled dynamic network that facilitates the diffusion of
potential health impacts.1 PIP across the interface. The increased amount of PIP in the TMC
solution results in a more homogeneous IP reaction, thus forming
3.3. Tuning reaction conditions of interfacial polymerization a polyamide network with a more uniform pore size distribution
Tuning reaction conditions of IP has been widely researched (Fig. 6C, right panel). Consequently, the SARIP membrane
since the origin of the TFC membranes. Previous efforts exhibited a sharper cut-off for ion separation than the conven-
through the control of monomer concentrations or adjustment tional IP membrane, potentially enabling precise separation in
of curing conditions of IP have resulted in numerous commercial desalination and water purification.
products.8,35 New strategies involve the use of new reaction
monomers and additives in IP for making high-performance 3.4. Incorporating emerging surface coating techniques into
TFC membranes. interfacial polymerization
Conventional IP is performed between a triacyl chloride The above-mentioned strategies involve the use of new chemicals
(particularly TMC) with a diamine—typically MPD for RO, (e.g., nanostrands, nanofillers, monomers, or additives) for tailoring
and piperazine (PIP) for NF. The reaction utilizes the high the IP process. Presently, polyamide chemistry made from MPD
reactivity between acyl chloride and amine groups, thereby and TMC monomers remains the gold standard for commercial RO
readily forming a dense selective layer at the organic–aqueous membranes. This fact has driven innovations in the manufacturing
solution interface. In essence, any pair of reactive chemicals, process of IP (Fig. 5D), while retaining the use of the conventional
which can be respectively dissolved in organic and aqueous MPD–TMC monomers—the pair that shows relevance and
phases, could be potential monomers for IP. For example, potential to the industrial fabrication process.
recent studies reported the use of hydroxyl-rich cyclodextrins As discussed above, the trans-interface diffusion of MPD
as alternative monomers of diamine to react with TMC.36 The molecules is the major factor limiting the homogeneous
porous cyclodextrin molecules endow the formed nanofilms formation of a polyamide network. Additionally, the fast
with unique Janus pathways (i.e., hydrophobic inner cavities kinetics of the MPD–TMC reaction lead to rapid formation of
and hydrophilic channels), thereby enabling precise molecular a polyamide network, as the solution at the interface readily
sieving in both organic solvent separation and desalination. reaches the gel point. This nascent polyamide gel provides
Notably, recent efforts have expanded the search of alternative additional resistance for the diffusion of the two monomers
monomers for IP to a broader range of chemicals as described during the reaction. Consequently, the conventional IP process
in these comprehensive reviews.37,38 produces selective polyamide layers with strong depth hetero-
Besides the use of new monomers, progress has also been geneity (i.e., variation in chemical structure from the surface to
made in adding various additives to the IP reaction the bottom) and rough surface morphology (i.e., characteristic
(Fig. 5C).39,40 Different from new monomers, these additives ridge-and-valley structure).41 To overcome these limitations,
do not become segments of the resulting polyamide polymer. molecular layer-by-layer (mLbL) deposition was proposed as
Instead, they influence the formed polyamide selective layer by an alternative IP process (Fig. 5D, lower panel).42 In this
inhibiting or facilitating the reaction between acyl chloride and technique, a pretreated substrate is sequentially immersed in
diamine at the interface. For example, in the conventional IP MPD and TMC; the immersion is repeated for several cycles to
process, the reaction predominantly occurs on the organic side obtain an integral polyamide film. Notably, after each soaking,
(i.e., TMC solution) of the interface, owing to the high diffusivity the substrate is thoroughly rinsed to leave only a thin layer of
of diamine monomers (e.g., MPD or PIP) from the aqueous the reaction solution on the surface. In this case, the mLbL
phase to the organic phase. Adding a macromolecule—polyvinyl technique builds the film with one molecular layer at a time via
alcohol (PVA)—to the aqueous phase effectively inhibits the instantaneous reaction of alternating MPD/TMC monomers,
diffusion of PIP, thus leading to a diffusion-driven Turing thereby facilitating complete reaction between the monomers to
instability at the interface (Fig. 5C, upper left panel).39 By form an integral polyamide network. Compared with membranes
adjusting the amount of the PVA additive, the surface made through conventional IP, the mLbL-polyamide film not only

6298 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev

has a smooth surface with enhanced fouling resistance, but regions targeting different applications (Fig. 7, shaded areas).
also exhibits increased NaCl rejection.42 Moreover, the mLbL The commercial membranes with excellent rejection of NaCl
technique enables fine-tuning of reaction conditions (e.g., monomer/ largely lay in the seawater desalination RO region (SWRO, grey
solvent type, concentration, and support materials), as well as area, RNaCl 4 99%), while the majority of the emerging TFC
nanoscale control of thickness, topology, and local chemical membranes are in the brackish water RO region (BWRO, yellow
composition of resulting polyamide films. Hence, the mLbL area, 99% 4 RNaCl 4 90%,) or nanofiltration region (NF, blue
technique holds promise as a versatile platform for revealing area, RNaCl o 90%). As seawater is the dominant feedwater
the reaction mechanisms of IP. source in the global share of desalination for water production
Compared with the mLbL technique, more precise control of (460%),6 substantial efforts are needed to improve the water–
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

reaction conditions could be achieved by integrating an electro- salt selectivity of the emerging TFC membranes, thereby
spray technique—an ionization method developed for mass improving their performance for SWRO uses.
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

spectroscopic analysis of polar biomolecules43—to the IP Notably, despite the performance gap between commercial and
process.44,45 In this process, MPD and TMC monomer solutions emerging TFC membranes made of the polyamide chemistries,
are respectively loaded into two syringes (Fig. 5D, upper panel). the latter largely outperform membranes made of new emerging
After applying a high voltage to the solution, the charged materials, such as nanotubes and nanosheets (purple hexagons),
solutions are ejected from the metallic needle and burst into in terms of water–salt selectivity. While the discussion on
fine sprays due to the repulsive Coulombic forces. When both
MPD and TMC sprays deposit on a substrate, the polymeriza-
tion reaction takes place to form a polyamide layer. In essence,
the dropwise deposition of monomer solutions confines the
MPD–TMC reaction to the interface of the microdroplets,
leading to the formation of a smooth polyamide film.44,45 This
process enables nanoscale control of the film thickness
(B4 nm) through simple manipulation of reaction conditions,
such as monomer concentration,44 spraying time,45 or use of
different supports.44 Notably, the formed polyamide film exhibits
a thin layer (B10 nm) and good NaCl rejection (495%),44
demonstrating the promise of this emerging IP process in making
a thin, defect-free polyamide film for desalination.

3.5. Performance summary of TFC membranes made through


emerging techniques
As discussed earlier in this Tutorial, TFC membranes are
constrained by the permeability–selectivity trade-off (Fig. 7).
The intrinsic properties of commercial TFC membranes made
by conventional IP (blue squares) lay beneath an upper-bound
line. Such a trade-off relationship also holds true for TFC mem-
branes fabricated through emerging IP strategies (red circles).
Although the two groups of membranes exhibit comparable water Fig. 7 Water permeability and water–salt selectivity of different types of
permeability values in the range of 1 to 5 L m2 h1 bar1, the desalination membranes. The blue squares represent the performance of
conventional-IP membranes substantially outperform the commercial TFC membranes made through conventional IP. These
emerging-IP membranes in water–salt selectivity. Such results values are adapted from ref. 22. The red circles represent lab-made TFC
membranes fabricated via some emerging IP strategies reviewed in this
are not surprising, as the commercial membranes are the pro-
Tutorial. The transport properties of some emerging membrane materials
ducts obtained after a long-term optimization process. Presently, are also included in the figure for comparison (purple hexagons). The
emerging IP strategies are still at the stage of lab-scale research, green and orange dashed lines correspond to NaCl rejections of 99% and
involving hand-cast procedures in making membranes. We expect 90%, respectively, at a water flux of 20 L m2 h1. The area under the upper
that some performance enhancement of the emerging-IP bound line is further divided into three regions based on the threshold
NaCl rejection of 99% and 90%. Most commercial TFC membranes have
membranes could be achieved by employing standard
rejections in the seawater reverse osmosis (SWRO) region (rejection
manufacturing processes. However, the potential and feasibility 4 99%), while the emerging TFC membranes have rejections in the
in scaling up these emerging IP strategies into real-world brackish water reverse osmosis (BWRO) region (90% o rejection o 99%)
industrial production processes should be carefully considered and the nanofiltration (NF) rejection (rejection o 90%). The separation
and evaluated. performance of the conventional TFC membranes is relatively closer to the
upper bound line, owing to the well-developed and optimized systems for
Notably, the low permselectivity of the emerging TFC
the conventional IP. In contrast, the data points of the emerging TFC
membranes was largely ascribed to the poor rejection of NaCl membranes are far below the upper bound line, suggesting that substantial
salt. For instance, based on projected NaCl rejection,46 the area efforts are needed to further optimize the fabrication conditions of the
beneath the upper bound line could be divided into several emerging IP strategies.

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6299
View Article Online

Chem Soc Rev Tutorial Review

desalination membranes made from emerging materials via observed on the top surface of TFC membranes. However,
non-IP methods is beyond the scope of this Tutorial, the low discerning the polyamide–polysulfone interface (i.e., the
performance of such membranes has been attributed to boundary between active and support layers) in such dark-
defects inherent in the top-down fabrication methods used.22 field images is difficult, because the polymeric materials with
Nevertheless, our analyses above demonstrate the promise of the dominant light elements (e.g., C, H, and O) on their backbones
polyamide chemistries as well as the IP-based processes for are nearly ‘‘transparent’’ to electrons. Such a technological gap
making high-performance desalination membranes. necessitates effective measures to enhance the contrast
difference (Fig. 8A), thereby illustrating fine structures of the
polyamide layer and the polyamide–polysulfone interface.
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

4. Emerging techniques in One simple strategy is to stain the polyamide layer with
characterization of TFC membranes heavy metal elements (such as osmium, ruthenium, uranyl, and
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

lead).48–50 In a general procedure, TFC membranes are


Insights gained through in-depth characterization of TFC immersed into a metal oxide solution, in which metal ions
membranes are important not only for understanding the readily approach and adhere to the charged regions of the
structure–property–performance relationship, but also for polyamide film. The stained membranes then undergo identical
future design and optimization of IP-based processes. Previous sample preparation processes (i.e., fixation, dehydration, and
efforts in probing TFC membrane structures largely relied on sectioning) and are imaged using electron microscopic techniques.
the use of conventional tools, such as spectroscopy [e.g., X-ray The metal stains absorb electrons or scatter part of the electron
photoelectron spectroscopy (XPS) and Fourier-transform beam which otherwise is projected onto the sample. As a result,
infrared (FITR) spectroscopy] and microscopy [e.g., scanning the stained polyamide layer is brighter than the unstained poly-
electron microscopy (SEM), transmission electron microscopy sulfone support in the dark-field images (Fig. 8a2). Previous studies
(TEM), and atomic Force Microscope (AFM)] techniques.47 utilized such enhanced contrast to map the distribution of
These tools have provided general structural information about functional groups in the polyamide layer.48,49 Recent
TFC membranes, like ridge-and-valley morphology, degree of efforts—through tailoring staining conditions, together with
crosslinking, and functional group distribution of the advances in a high-resolution electron microscope—have
polyamide layer, facilitating the understanding of the TFC enabled a closer look at the inner structure of the polyamide
membranes. layer.50 For example, precipitation of the metal stains on the outer
As discussed above, IP is a rapid self-inhibiting process that surface of the polyamide clearly illustrates the nodular shape of
enables the formation of a thin, defect-free polyamide film. the polyamide (Fig. 8a2). Moreover, the distribution of aqueous
These features of IP, in turn, pose challenges for nanoscale stain tracers reflects actual transport pathways in the polyamide
characterization of TFC membranes. For instance, when exploring layer, thereby revealing the ‘‘channel’’ structure for water transport
IP reaction mechanisms, the fast reaction kinetics between amine within the polyamide layer (inset of Fig. 8a2).50 Although such
and acyl chloride limit the direct observation of polyamide ‘‘channels’’ were also observed in non-stained samples (white
network formation during the IP reaction using conventional arrows in Fig. 8a1), staining substantially enhanced the contrast
characterization tools. Moreover, the formed thin polyamide film for visualizing such fine structures.
(B100 nm) only accounts for B1% of the total thickness of a Elemental mapping is another strategy for enhancing inter-
typical TFC membrane (B100 mm for the underlying polysulfone facial contrast.41 In particular, the polyamide active layer is rich
support and PET fabric layers). When characterizing the in nitrogen while the underlying polysulfone support layer
polyamide layer via conventional spectroscopic techniques, the contains sulfur but no nitrogen. This intrinsic elemental contrast
underlying support materials generate substantial amounts of allows researchers to perform an analysis using scanning
background noise that interferes with the characteristic transmission electron microscopy (STEM) coupled with energy-
information of polyamide, hampering a closer look at the dispersive X-ray spectroscopy (EDX), for elucidating the nanoscale
polyamide layer alone. Consequently, along with the progress in structure and chemical composition.41 In the mapping image
developing emerging IP strategies, recent years have also (Fig. 8a3), fine structures at the polyamide–polysulfone interface
witnessed substantial efforts in exploring new techniques for could be observed. For example, the presence of cavities inside the
TFC membrane characterization. These newly developed techniques ridge structure of both membranes (white arrows) suggests depth
could be classified into two categories: direct imaging techniques heterogenicity of the polyamide film. Additionally, the polyamide
and indirect non-imaging techniques. An overview of these layer comprises a continuous 50 nm-thick polyamide base layer
techniques is presented below, and the pros and cons of each (yellow arrow and bracket), from which the ridge-and-valley
technique are reviewed and discussed. structure extends outward.
Direct imaging of polyamide internal microstructure could
4.1. Direct imaging techniques also be achieved through the isolation of the polyamide layer,
Direct imaging techniques aim to provide visual structural followed by 3D tomography visualization (Fig. 8B).51 In the
information of the polyamide layer by using electron micro- isolation step, the TFC membrane is immersed in a polar
scopic tools (Fig. 8). In a regular cross-section TEM image solvent (e.g., DMF, NMP, and THF) to dissolve the underlying
(Fig. 8a1), a characteristic ridge-and-valley structure could be polysulfone layer, while the structure of the insoluble

6300 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev


This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

Fig. 8 Direct imaging techniques for the characterization of TFC membranes. (A) Enhancement of interfacial contrast for the direct visualization of
polyamide–polysulfone interfacial structure. Comparison of scanning transmission electron microscopy (STEM) cross-section images of a non-stained
(a1) and a stained (a2) TFC membrane. Metal stains bound to the polyamide active layer result in convoluted regions (white arrowhead), enabling
differentiation of the interface and observation of the polyamide internal structure. The inset shows higher magnification of a region of polyamide layer
with empty (white arrow) and filled (yellow arrow) channels. Scanning electron microscopy images are adapted with permission from ref. 50, Elsevier. (a3)
Elemental mapping also presents a clear interfacial structure of a TFC membrane. This technique utilizes the elemental contrast between the polyamide
(abundant in nitrogen) and polysulfone (abundant in sulfur) layers. ‘‘N’’ (magenta) denotes nitrogen and ‘‘S’’ (green) denotes sulfur; the white arrows
indicate the cavities inside the ridge structure, while the yellow arrow and bracket indicate the dense polyamide layer. Scanning electron microscopy
image is adapted with permission from ref. 41, American Chemical Society. (B) Isolation of the polyamide layer for 3D tomographic characterization. (b1)
Schematic illustration of tomography techniques. A range of 2D images are collected using an advanced electron microscope by tilting the polyamide
film sample to many different angles. The 3D structure of the sample is then reconstructed from the images, thereby enabling further analysis of its
structural characteristics. (b2) 3D visualization of the spatial distribution of pore structure of various sizes within the interior of a polyamide film isolated
from a commercial TFC membrane. The pores are classified based on their respective internal sizes: singe voids have a size larger than 30 nm (green)
while the spheres have a size smaller than 30 nm (red) or 15 nm (yellow). Some single voids are connected to form some giant pores (blue). Figure is
adapted with permission from ref. 51, Elsevier. (b3) Tomographs of two commercial TFC membranes for SWRO and BWRO applications. The intensity of
the polyamide layers is displayed as a heat map, where red corresponds to higher intensity. Figure is adapted with permission from ref. 52, United States
National Academy of Sciences.

polyamide network is largely retained. The isolated polyamide providing high-quality images of soft materials for nanoscale
film is then imaged by TEM 3D tomography. By rotating the analysis. Moreover, STEM tomography operated in high-angle
sample around the tilt axis, a series of 2D projection images are annular dark-field (HAADF) mode can provide additional
acquired at different tilt angles (normally within a range of information about polymeric structures, such as mass,
791 to +791). After alignment and processing, the tilt series are thickness, and atomic number.52,53 For instance, in heat
reconstructed into a 3D model of the isolated polyamide film, maps of the STEM images of two commercial TFC
thereby providing information on its internal structure.51 For membranes (Fig. 8b3), the membrane for seawater RO
example, in a tomogram of a commercial TFC membrane (SWRO, upper panel) exhibits a much denser structure (red
(Fig. 8b2), internal voids with various sizes and shapes were area) than that of the membrane for brackish water RO (BWRO,
clearly illustrated.51 These voids, which govern the pathways of lower panel).52 The denser structure is likely the actual selective
flow and diffusion across the polyamide layer, were either barrier of polyamide, which is responsible for the higher salt
connected to the outside of the polyamide film or rejection and lower water permeance of the SWRO membrane
encapsulated within the polyamide matrix. compared to the BWRO.52,53
Following the progress in TEM tomography, scanning TEM
(STEM) has also been introduced for imaging polyamide layers 4.2. Indirect non-imaging techniques
(Fig. 8b1).52,53 Unlike the use of a parallel, coherent beam of Parallel to efforts in direct imaging techniques, progress has
electrons for imaging in conventional TEM, STEM technique also been made in using non-imaging techniques for TFC
scans a focused beam of electrons across a sample. Such membrane characterization. These techniques, coupled with
dynamic focus effectively reduces the defocus variation in spectroscopic or scattering tools, shed light on chemical
TEM imaging of a thick sample at a high-tilt angle,52 thereby composition, nanopore size distribution, and polymeric chain

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6301
View Article Online

Chem Soc Rev Tutorial Review

structure within the polyamide layer, thereby providing layer, where each positron combines with an electron to form an
additional valuable information for revealing the membrane’s electron–positronium (Ps) ion. These Ps ions tend to locate in
internal structure. electron-deficient regions, such as free volume and pores, and
As discussed earlier, conventional spectroscopic methods undergo a slow self-annihilation process with a long lifetime. In
provide limited information about the polyamide internal electron-rich regions that were occupied by the polyamide
structure. Researchers sought to develop effective characterization matrix, the Ps ions would interact with surrounding molecularly
strategies to fully explore the capability of the spectroscopic bound electrons to induce a rapid annihilation process with
techniques. An earlier example was impregnating the polyamide a short lifetime. By correlating the lifetime of the injected
layer with silver nitrate solution, in which silver ions selectively positrons with the sample void size, PALS could reveal charac-
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

bind to negatively charged carboxyl groups at a 1 : 1 ratio.54 teristics of pore structure, such as size, distribution, and film
The impregnated polyamide is then characterized by Rutherford thickness. For example, using this technique, researchers were
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

backscattering spectrometry (RBS), which detects the amount of able to reveal the presence of two types of pores within the
the bound silver probe for determining the volume density of polyamide layer of an RO membrane, i.e., smaller intrachain
carboxyl groups. Using this technique, other useful information, network pores (B0.45 nm) and larger interchain aggregate pores
such as ionization behaviors and ion accessibility, could also be (B0.8 nm).57 This sub-nanometer insight has enabled substantial
revealed by adjusting characterization conditions (e.g., impregnation progress in elucidating the structure–performance relationship of
pH or ion probe types).54 However, this silver-RBS method has not TFC membranes, as well as in understanding the reaction
been widely used in membrane characterization, largely owing to mechanisms on IP. Readers are referred to a recent comprehensive
the inaccessibility of the specialized RBS instrument. In a later review for a detailed introduction of PALS in membrane
effort,55 researchers modified the silver-binding procedure used in characterization.58
RBS by including a low-pH elution step, which allowed the use of a Besides advances in spectroscopic techniques, recent years
more accessible spectroscopic technique—inductively coupled have also witnessed increasing interest in using scattering
plasma mass spectrometry—to quantify the eluted silver concen- techniques for TFC membrane characterization (Fig. 9).59–61
tration for determining carboxyl density of the polyamide layer. During the measurements, the incident sources (e.g., X-ray or
In addition to the volume density of functional groups, neutron) are scattered by the electrons (X-ray) or the nuclei
spectroscopic techniques could also be combined with incident (neutron) to generate corresponding signals (Fig. 9A). Generally,
high beams for revealing the depth profile of the polyamide the interference of scattered sources from neighboring planes
layer.41,56 For example, XPS—a technique that only probes the could be described by Bragg’s law: nl = 2d sin y, where y is the
very top surface of the polyamide layer (o5 nm)—could be scattering angle, d is the distance between the atomic planes
paired with C60+ cluster ion sputtering platform.41 In this (d-spacing), l is the wavelength of the incident source, and n is
combined technique, the polyamide side of TFC membranes an integer. As the scattering vector, q, is related to y by the
is exposed to a high beam generated from C60+ bombardment, relation q = 4p sin y/l, Bragg’s law could be written as d = 2p/q. In
enabling controllable sputtering of the polyamide layer. At principle, Bragg’s law is applicable to crystalline materials with
different sputtering times, XPS elemental scans are performed repeating discrete reflections. However, the inverse relationship
on the sputtered spot in order to reveal the atomic composition of the actual length scale, d, and the reciprocal space, q, could
and chemical state of the organic thin films with depth.41 In a still be used in the interpretation of the interchain distance in
similar technique called elastic recoil detection (ERD), a amorphous materials, such as highly crosslinked polyamide. For
membrane sample is bombarded with a high-energy iodine example, wide-angle X-ray scattering (WAXS) generally collects
beam, causing target atoms (e.g., C, N, O, S, and Cl) to be scattered signals with q 4 0.5 Å1, corresponding to the atomic
scattered out of the sample.56 Analysis of the scattered signal or molecular structures with d = 0.1–10 Å, while small-angle X-ray
not only identifies the recoil ion species according to their scattering (SAXS) or small-angle neutron scattering (SANS)
respective nuclear charge or mass, but also detects the energy collect scattered signals with q range of 0.005–0.5 Å1 for
of the recoil ions.56 This latter information is then used to revealing mesoscale structure (i.e., d = 10–1000 Å) from the
determine the energy loss between the projectile and the recoil aggregates or clusters of atoms and molecules.
ion within the sample matrix, thereby enabling analysis of Two types of polyamide samples are commonly used in
elemental depth profiling of the membrane sample. Notably, scattering experiments: polyamide film and bulk dispersion.
the depth profiles extracted from these spectroscopic tools (i.e., In film characterization mode (Fig. 9B), incident X-ray is shed
XPS-C60+ and ERD) revealed the presence of a heterogeneous on a polyamide film at a relatively small angle, and scattered
layer that contains both polyamide and polysulfone signals are then detected by a 2D detector.59 As the film is
signatures.41,56 Such a structure implies that the polyamide horizontally positioned on a sample stage, analysis of scattering
penetrates the pores of polysulfone during IP reaction, anchoring patterns could provide information on both molecular structure
to the support to form a stable TFC structure. and preferential orientation (either in-plane or out-of-plane) of
Positron annihilation lifetime spectroscopy (PALS) is the polyamide network.59 For example, an arc-like meridional
another technique capable of probing the internal pore structure scattering feature in a WAXS pattern (Fig. 9C), where intense
of the polyamide layer on TFC membranes.57 The analysis is scattering is preferentially aligned along the vertical qz direction,
conducted by injecting low-energy positrons into the polyamide implies the in-plane orientation of aromatic stacking in the

6302 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev


This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

Fig. 9 Indirect scattering techniques for the characterization of TFC membranes. (A) Interaction of atomic matter with scattering sources to generate
scattering signals. X-rays interact with electrons in the material via an electromagnetic force, while neutrons interact with the nucleus via the very short-
range strong nuclear force. (B) Schematic illustration of the experimental setup used for characterizing a polyamide film using wide-angle X-ray
scattering (WAXS). A representative anisotropic pattern on a 2D WAXS detector provides information on the molecular structure and preferential
orientations in the polyamide film. (C) A representative WAXS pattern of a polyamide film. The horizontal (qr) and vertical (qz) vectors correspond to the
surface parallel and normal scattering wave vectors, respectively. Dashed arcs are scattering vectors of 1.22, 1.54, and 1.79 Å1, corresponding to the
spacing of 5.2, 4.1, and 3.5 Å, respectively. Figure is adapted with permission from ref. 59, American Chemical Society. (D) Different stacking behaviors of
aromatic rings within a polyamide layer. The green dashed lines indicate the distance of 3.5 Å for the aromatic cores stacked in p–p (parallel) structure,
while the red dash lines indicate the distance of 5.2 Å for the aromatic cores stacked in T-shaped (orthogonal) structure. (E) Schematic illustration of
small-angle X-ray scattering (SAXS) platform used for characterizing polyamide bulk dispersion. Solid polyamide film is ground into small flakes and
dispersed in water as a colloidal-like suspension, thus resulting in a pattern of isotropic rings on a 2D SAXS detector. (F) Schematic diagram of 1D
scattering intensity profile of polyamide by integrating a 2D SAXS pattern. The profile can be divided into three regions for respective analysis. In the
Guinier region, scattering intensity, I, is proportional to the concentration of the system and thus independent of q. In the mass fractal (I p q2–3) and the
Porod region (I p q3–4), the correlation of I with q can provide the structural information of large clusters and small primary units of polyamide chain,
respectively.

polyamide film.59 Further analysis of this WAXS pattern by Prior to the scattering experiments, freshly-prepared polyamide
plotting the scattered intensity versus scattering vector, q, films are either loaded into a small quartz capillary,61 or ground
suggests the presence of two peaks centered at 1.79 Å1 and into small flakes and dispersed in water as a colloidal-like
1.22 Å1. According to Bragg’s law (d = 2p/q), these two peaks suspension.60 Integration of 2D scattering pattern results in a
correspond to molecular spacings of 3.5 and 5.2 Å, respectively. 1D scattering profile (schematic illustrated in Fig. 9F), which
The 3.5 Å spacing is consistent with ‘‘p–p’’ stacking of aromatic could be divided into three regions for respective analysis.61
cores (green arrows in Fig. 9D), whereas the 5.2 Å spacing is In the Guinier region (extremely low q), scattering intensity, I,
consistent with the perpendicular packing of aromatic cores in is proportional to the concentration of the system and thus
the ‘‘T-shaped’’ configuration (red arrows in Fig. 9D).59 Based on independent of q. At a higher q, the correlation of I with q could
such WAXS results, researchers revealed the important role of provide information on polyamide chain structure, including
reaction kinetics and post-treatment procedures in affecting the larger clusters or smaller primary units. For example, in the
stacking behaviors of aromatic rings. For example, the emerging Porod region (I p q3–4), the SAXS pattern of an RO polyamide
mLbL IP strategy with well-controlled MPD–TMC reaction layer reveals that the primary globular units have a radius of
leads to the generation of parallel ‘‘p–p’’ stacking, while the B130 Å, containing B1500 phenylenediamine trimesamide
conventional IP processes with uncontrolled MPD–TMC reaction units.61 In comparison, the primary unit of an NF polypiperazine
give rise to ‘‘T-shaped’’ configurations. It is surmised that the layer is slightly smaller, having radii of B120 Å with B1400
‘‘T-shaped’’ stacking with loose structure is more favorable for piperazine trimesamide units.61 Additionally, a steeper slope for
water diffusion, thereby governing the mass transport of the NF membranes (log I/log q = 3.8) than that for the RO
polyamide films. membranes (log I/log q = 3.4) suggests that the globular units
Besides the molecular-level insights from the WAXS in the NF membranes have a relatively higher degree of
measurement of polyamide films, SAXS (or SANS) measurements compaction than those in the RO membranes.61 Notably, the
were used to characterize polyamide dispersions for revealing aggregation of these primary units leads to the formation of
structural information at the polymer-chain level (Fig. 9E).60,61 larger clusters, which could also be characterized by the SAXS

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6303
View Article Online

Chem Soc Rev Tutorial Review

profiles in the mass fractal regions (I p q2–3). For instance, the IP reaction involves the generation of heterogeneous
nodular clusters of the NF polypiperazine network have a size of environments with both liquid solution and solid polyamide,
1900–3000 Å, which is six times larger than that of the RO adding further constraints to the system for in situ characteriza-
polyamide network. Taken together, these insights from the tion. For the resulting membranes, existing characterization is
scattering studies suggest that the RO polyamide layer has larger largely conducted under ambient conditions, while operando
primary units, lower chain compaction, and smaller size characterization of TFC membranes under conditions relevant
clusters than those of the NF polypiperazine layer.60,61 Such to real-world operation (e.g., high pressure, crossflow, and
differences in the polymer network structures could be related salinity) is hampered by numerous technical obstacles. For
to their desalination performance, thereby facilitating the instance, electron microscopic tools require stringent sample
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

understanding of the IP reaction for guiding the design of pretreatment (e.g., drying, sectioning, or coating), thereby
high-performance TFC membranes. limiting their uses for operando characterization. Although
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

X-ray (or neutron) scattering techniques do not require sample


pretreatment, incorporating custom-built setups, which can
5. Conclusion and outlook mimic the operating conditions, into existing facilities is extremely
difficult. Given the importance of in situ or operando information
In this Tutorial, we presented a brief history of the development in delineating reaction mechanisms of IP and structure–property–
of desalination membranes, discussed the current challenges performance relationship of TFC membranes, increasing efforts
in the application of conventional TFC membranes, and should be devoted to developing relevant characterization tools
reviewed emerging progress in IP-based fabrication processes and platforms in future research.
and TFC membrane characterization techniques. Despite In addition to the efforts in developing experimental measures
extensive research efforts in the past decade, the industrial IP with the required spatial and temporal resolution, molecular
process and its product polyamide layer are largely unchanged simulations can be a versatile tool to gain insights into the IP
since the invention of TFC membranes in the late 1970s, reaction kinetics and polyamide formation mechanisms. For
suggesting knowledge gaps between lab research and real-world example, a recent study using molecular dynamic simulation
applications. Herein, we further discuss key limiting factors of the predicted the presence of ‘‘p–p’’ and ‘‘T-shaped’’ stacking for
IP process and propose potential research directions for fabricating the aromatic rings of the polyamide structure,65 thereby guiding
next-generation TFC desalination membranes. the selection and design of pertinent characterization tools in
Development of new IP strategies has been an active area of revealing such microstructures (see the above section about X-ray
research in the past decade, as reviewed in this Tutorial. scattering techniques).59 In another simulation study, researchers
Presently, obtaining a polyamide layer with a smooth surface suggested a reaction-aggregation process of IP in forming small
for fouling control is possible using some emerging polyamide oligomers, which then aggregated to yield large, solid
approaches, such as electrospray or constructing sacrificial polyamide clusters to limit further reaction of IP.66 Notably, their
interlayers. However, effective strategies for making membranes modeling results were also validated by an experimental study,67
with ultrahigh water–salt selectivity are still lacking, mainly demonstrating the promise of the developed model in studying
owing to the challenges in obtaining a thin, defect-free poly- the complex IP reaction kinetics. Despite such progress, current
amide film after adjusting reaction conditions of IP. Given the efforts in the simulation of IP are still in their infancy, largely
importance of enhanced water–salt selectivity in improving relying on molecular dynamic simulations.65,66,68 With recent
water quality and reducing the cost of desalination, future efforts advances in computation, molecular simulations can play more
in developing new IP strategies would require effective measures important roles in understanding IP reaction mechanisms and
of molecular-level manipulation of polyamide microstructures guiding TFC membrane design. For instance, the utilization of
where water and solute transport takes place. For example, supercomputers can enable more complicated simulations of the
controlling reaction or posttreatment conditions of IP resulted IP reaction by taking more reaction factors into consideration.
in different molecular motifs of polyamide, potentially affecting The development of machine learning or artificial intelligence can
mass transport in TFC membranes.59 Such insights may lead to also relieve researchers of some tedious, time-consuming experi-
new approaches for making high-performance membranes. ments in finding the best reaction conditions.
Notably, advances in materials science related to thin-film Last, in addition to research efforts in academia, the design
fabrication techniques (e.g., self-assembly,62 atomic layer of IP processes for making next-generation TFC membranes
deposition,63 and surface-initiated polymerization64) can provide would also require collaborative efforts with the industry. After
new approaches for the design of novel IP strategies. long-term optimization of the conventional IP process,
The use of advanced characterization techniques has membrane industries have established an inventory of
enabled substantial progress in the mechanistic understanding empirical protocols to making high-performance TFC
of the IP reaction and membrane structure. However, techniques membranes for different applications, including seawater
for in situ characterization of IP reaction kinetics, as well as desalination, wastewater reclamation, or even tap water
operando characterization of TFC membranes, are still lacking. softening. However, owing to the lack of interactions between
As discussed in this Tutorial, the fast kinetics of the IP reaction academia and industry, a large portion of research efforts in
limit direct observation of polyamide formation. Moreover, the academia still focus on tuning minor conditions of IP to

6304 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev

achieve marginal gains in membrane performance, which may 10 H. B. Park, J. Kamcev, L. M. Robeson, M. Elimelech and
have been realized by industry a long time ago. To avoid such B. D. Freeman, Maximizing the right stuff: The trade-off
superficial repetition in future research, academic researchers between membrane permeability and selectivity, Science,
should put more emphasis on revealing IP reaction mechanisms 2017, 356, eaab0530.
or demonstrating proof-of-concept of new IP strategies, while 11 J. R. Werber, A. Deshmukh and M. Elimelech, The Critical
seeking collaborations with industrial partners in scaling up lab Need for Increased Selectivity, Not Increased Water Perme-
results into commercial products. Such collaborative experiences ability, for Desalination Membranes, Environ. Sci. Technol.
will enable academic researchers to understand not only the Lett., 2016, 3, 112–120.
design factors for high-performance membranes, but also the 12 D. Cohen-Tanugi, R. K. McGovern, S. H. Dave, J. H. Lienhard
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

factors that afford scalable production processes. and J. C. Grossman, Quantifying the potential of ultra-
permeable membranes for water desalination, Energy
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

Environ. Sci., 2014, 7, 1134–1141.


Conflicts of interest 13 X. Zhou, Z. Wang, R. Epsztein, C. Zhan, W. Li, J. D. Fortner,
T. A. Pham, J.-H. Kim and M. Elimelech, Intrapore energy
There are no conflicts to declare. barriers govern ion transport and selectivity of desalination
membranes, Sci. Adv., 2020, 6, eabd9045.
14 R. Zhang, Y. Liu, M. He, Y. Su, X. Zhao, M. Elimelech and
Acknowledgements Z. Jiang, Antifouling membranes for sustainable water
purification: strategies and mechanisms, Chem. Soc. Rev.,
We acknowledge the financial support received from the NSF
2016, 45, 5888–5924.
Nanosystems Engineering Research Center for Nanotechnology-
15 D. Rana and T. Matsuura, Surface Modifications for Anti-
Enabled Water Treatment (EEC-1449500). This work was partly
fouling Membranes, Chem. Rev., 2010, 110, 2448–2471.
supported by the Fundamental Research Funds for the Central
16 G.-D. Kang and Y.-M. Cao, Development of antifouling
Universities. We also acknowledge fruitful discussions with Wen
reverse osmosis membranes for water treatment: A review,
Ma, Jay Werber, Akshay Deshmukh, and Rhea Verbeke at Yale
Water Res., 2012, 46, 584–600.
University.
17 J. Glater, S. K. Hong and M. Elimelech, The Search for a
Chlorine-Resistant Reverse-Osmosis Membrane, Desalina-
tion, 1994, 95, 325–345.
References
18 R. Verbeke, V. Gómez and I. F. J. Vankelecom, Chlorine-
1 M. S. Mauter, I. Zucker, F. Perreault, J. R. Werber, J.-H. Kim resistance of reverse osmosis (RO) polyamide membranes,
and M. Elimelech, The role of nanotechnology in tackling Prog. Polym. Sci., 2017, 72, 1–15.
global water challenges, Nat. Sustain., 2018, 1, 166–175. 19 Y. Yao, P. Zhang, C. Jiang, R. M. DuChanois, X. Zhang and
2 M. A. Shannon, P. W. Bohn, M. Elimelech, J. G. Georgiadis, M. Elimelech, High performance polyester reverse osmosis
B. J. Marinas and A. M. Mayes, Science and technology for desalination membrane with chlorine resistance, Nat. Sus-
water purification in the coming decades, Nature, 2008, 452, tain., 2021, 4, 138–146.
301–310. 20 S. Karan, Z. Jiang and A. G. Livingston, Sub–10 nm
3 M. Elimelech and W. A. Phillip, The Future of Seawater polyamide nanofilms with ultrafast solvent transport for
Desalination: Energy, Technology, and the Environment, molecular separation, Science, 2015, 348, 1347–1351.
Science, 2011, 333, 712–717. 21 X.-H. Ma, Z.-K. Yao, Z. Yang, H. Guo, Z.-L. Xu, C. Y. Tang and
4 R. Weaver, M. Howells, Y. Yang, S. Lennox and H. Brown, M. Elimelech, Nanofoaming of Polyamide Desalination
IDA Water Security Handbook 2019–2020, 2020. Membranes To Tune Permeability and Selectivity, Environ.
5 A. Deshmukh, C. Boo, V. Karanikola, S. Lin, A. P. Straub, Sci. Technol. Lett., 2018, 5, 123–130.
T. Tong, D. M. Warsinger and M. Elimelech, Membrane 22 J. R. Werber, C. O. Osuji and M. Elimelech, Materials for
distillation at the water-energy nexus: limits, opportunities, next-generation desalination and water purification
and challenges, Energy Environ. Sci., 2018, 11, 1177–1196. membranes, Nat. Rev. Mater., 2016, 1, 16018.
6 E. Jones, M. Qadir, M. T. H. van Vliet, V. Smakhtin and 23 Z. Wang, Z. Wang, S. Lin, H. Jin, S. Gao, Y. Zhu and J. Jin,
S.-M. Kang, The state of desalination and brine production: Nanoparticle-templated nanofiltration membranes for
A global outlook, Sci. Total Environ., 2019, 657, 1343–1356. ultrahigh performance desalination, Nat. Commun., 2018,
7 J. Glater, The early history of reverse osmosis membrane 9, 2004.
development, Desalination, 1998, 117, 297–309. 24 Z. Yang, P.-F. Sun, X. Li, B. Gan, L. Wang, X. Song, H.-D.
8 R. W. Baker, Membrane Technology and Applications, Wiley, Park and C. Y. Tang, A Critical Review on Thin-Film
New York, 2nd edn, 2004. Nanocomposite Membranes with Interlayered Structure:
9 G. M. Geise, H. B. Park, A. C. Sagle, B. D. Freeman and Mechanisms, Recent Developments, and Environmental
J. E. McGrath, Water permeability and water/salt selectivity Applications, Environ. Sci. Technol., 2020, 54, 15563–15583.
tradeoff in polymers for desalination, J. Membr. Sci., 2011, 25 S.-J. Park, W. Choi, S.-E. Nam, S. Hong, J. S. Lee and J.-H.
369, 130–138. Lee, Fabrication of polyamide thin film composite reverse

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6305
View Article Online

Chem Soc Rev Tutorial Review

osmosis membranes via support-free interfacial polymeriza- 41 X. Lu, S. Nejati, Y. Choo, C. O. Osuji, J. Ma and
tion, J. Membr. Sci., 2017, 526, 52–59. M. Elimelech, Elements Provide a Clue: Nanoscale Charac-
26 Z. Jiang, S. Karan and A. G. Livingston, Water Transport terization of Thin-Film Composite Polyamide Membranes,
through Ultrathin Polyamide Nanofilms Used for Reverse ACS Appl. Mater. Interfaces, 2015, 7, 16917–16922.
Osmosis, Adv. Mater., 2018, 30, 1705973. 42 G. Joung-Eun, L. Seunghye, C. M. Stafford, L. J. Suk,
27 M. M. Pendergast and E. M. V. Hoek, A review of water C. Wansuk, K. Bo-Young, B. Kyung-Youl, E. P. Chan,
treatment membrane nanotechnologies, Energy Environ. C. J. Young, B. Joona and L. Jung-Hyun, Molecular
Sci., 2011, 4, 1946–1971. Layer-by-Layer Assembled Thin-Film Composite Mem-
28 Y. Zhao, C. Qiu, X. Li, A. Vararattanavech, W. Shen, J. Torres, branes for Water Desalination, Adv. Mater., 2013, 25,
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

C. Hélix-Nielsen, R. Wang, X. Hu, A. G. Fane and C. Y. Tang, 4778–4782.


Synthesis of robust and high-performance aquaporin-based 43 J. Fenn, M. Mann, C. Meng, S. Wong and C. Whitehouse,
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

biomimetic membranes by interfacial polymerization- Electrospray ionization for mass spectrometry of large bio-
membrane preparation and RO performance characteriza- molecules, Science, 1989, 246, 64–71.
tion, J. Membr. Sci., 2012, 423-424, 422–428. 44 M. R. Chowdhury, J. Steffes, B. D. Huey and
29 X. Li, S. Chou, R. Wang, L. Shi, W. Fang, G. Chaitra, J. R. McCutcheon, 3D printed polyamide membranes for
C. Y. Tang, J. Torres, X. Hu and A. G. Fane, Nature gives desalination, Science, 2018, 361, 682–686.
the best solution for desalination: Aquaporin-based hollow 45 X.-H. Ma, Z. Yang, Z.-K. Yao, H. Guo, Z.-L. Xu and C. Y. Tang,
fiber composite membrane with superior performance, Interfacial Polymerization with Electrosprayed Microdro-
J. Membr. Sci., 2015, 494, 68–77. plets: Toward Controllable and Ultrathin Polyamide Mem-
30 C. J. Porter, J. R. Werber, M. Zhong, C. J. Wilson and branes, Environ. Sci. Technol. Lett., 2018, 5, 117–122.
M. Elimelech, Pathways and Challenges for Biomimetic 46 Z. Yang, H. Guo and C. Y. Tang, The upper bound of thin-
Desalination Membranes with Sub-Nanometer Channels, film composite (TFC) polyamide membranes for desalina-
ACS Nano, 2020, 14, 10894–10916. tion, J. Membr. Sci., 2019, 590, 117297.
31 E. M. V. Hoek and V. V. Tarabara, Encyclopedia of membrane 47 P. S. Singh, A. P. Rao, P. Ray, A. Bhattacharya, K. Singh,
science and technology, Wiley Online Library, 2013. N. K. Saha and A. V. R. Reddy, Techniques for characteriza-
32 Z. Yang, H. Guo, Z.-K. Yao, Y. Mei and C. Y. Tang, Hydro- tion of polyamide thin film composite membranes, Desali-
philic Silver Nanoparticles Induce Selective Nanochannels nation, 2011, 282, 78–86.
in Thin Film Nanocomposite Polyamide Membranes, 48 V. Freger, A. Bottino, G. Capannelli, M. Perry, V. Gitis and
Environ. Sci. Technol., 2019, 53, 5301–5308. S. Belfer, Characterization of novel acid-stable NF mem-
33 J. Yin, Z. Yang, C. Y. Tang and B. Deng, Probing the branes before and after exposure to acid using ATR-FTIR,
Contributions of Interior and Exterior Channels of TEM and AFM, J. Membr. Sci., 2005, 256, 134–142.
Nanofillers toward the Enhanced Separation Performance 49 C. Y. Tang, Y.-N. Kwon and J. O. Leckie, Probing the nano-
of a Thin-Film Nanocomposite Reverse Osmosis Membrane, and micro-scales of reverse osmosis membranes—A com-
Environ. Sci. Technol. Lett., 2020, 7, 766–772. prehensive characterization of physiochemical properties of
34 M. Mulder, Basic Principles of Membrane Technology, Kluwer uncoated and coated membranes by XPS, TEM, ATR-FTIR,
Academic Pulisher, The Netherlands, 2nd edn, 1996. and streaming potential measurements, J. Membr. Sci.,
35 R. J. Petersen, Composite Reverse-Osmosis and Nanofiltra- 2007, 287, 146–156.
tion Membranes, J. Membr. Sci., 1993, 83, 81–150. 50 M. M. Kłosowski, C. M. McGilvery, Y. Li, P. Abellan,
36 J. Liu, D. Hua, Y. Zhang, S. Japip and T.-S. Chung, Precise Q. Ramasse, J. T. Cabral, A. G. Livingston and A. E. Porter,
Molecular Sieving Architectures with Janus Pathways for Micro-to nano-scale characterisation of polyamide struc-
Both Polar and Nonpolar Molecules, Adv. Mater., 2018, tures of the SW30HR RO membrane using advanced elec-
30, 1705933. tron microscopy and stain tracers, J. Membr. Sci., 2016, 520,
37 M. J. T. Raaijmakers and N. E. Benes, Current trends in 465–476.
interfacial polymerization chemistry, Prog. Polym. Sci., 2016, 51 F. Pacheco, R. Sougrat, M. Reinhard, J. O. Leckie and
63, 86–142. I. Pinnau, 3D visualization of the internal nanostructure
38 A. F. Ismail, M. Padaki, N. Hilal, T. Matsuura and W. J. Lau, of polyamide thin films in RO membranes, J. Membr. Sci.,
Thin film composite membrane—Recent development and 2016, 501, 33–44.
future potential, Desalination, 2015, 356, 140–148. 52 T. E. Culp, Y.-X. Shen, M. Geitner, M. Paul, A. Roy,
39 Z. Tan, S. Chen, X. Peng, L. Zhang and C. Gao, Polyamide M. J. Behr, S. Rosenberg, J. Gu, M. Kumar and
membranes with nanoscale Turing structures for water E. D. Gomez, Electron tomography reveals details of the
purification, Science, 2018, 360, 518–521. internal microstructure of desalination membranes, Proc.
40 Y. Liang, Y. Zhu, C. Liu, K.-R. Lee, W.-S. Hung, Z. Wang, Natl. Acad. Sci. U. S. A., 2018, 115, 8694–8699.
Y. Li, M. Elimelech, J. Jin and S. Lin, Polyamide nanofiltra- 53 T. E. Culp, B. Khara, K. P. Brickey, M. Geitner, T. J. Zimudzi,
tion membrane with highly uniform sub-nanometre pores J. D. Wilbur, S. D. Jons, A. Roy, M. Paul,
for sub-1 Å precision separation, Nat. Commun., 2020, B. Ganapathysubramanian, A. L. Zydney, M. Kumar and
11, 2015. E. D. Gomez, Nanoscale control of internal inhomogeneity

6306 | Chem. Soc. Rev., 2021, 50, 6290–6307 This journal is © The Royal Society of Chemistry 2021
View Article Online

Tutorial Review Chem Soc Rev

enhances water transport in desalination membranes, 61 P. S. Singh, P. Ray, Z. Xie and M. Hoang, Synchrotron SAXS
Science, 2021, 371, 72–75. to probe cross-linked network of polyamide ‘reverse osmo-
54 O. Coronell, B. J. Marinas, X. J. Zhang and D. G. Cahill, sis’ and ‘nanofiltration’ membranes, J. Membr. Sci., 2012,
Quantification of functional groups and modeling of their 421-422, 51–59.
ionization behavior in the active layer of FT30 reverse 62 X. Feng, Q. Imran, Y. Zhang, L. Sixdenier, X. Lu,
osmosis membrane, Environ. Sci. Technol., 2008, 42, G. Kaufman, U. Gabinet, K. Kawabata, M. Elimelech and
5260–5266. C. O. Osuji, Precise nanofiltration in a fouling-resistant self-
55 D. Chen, J. R. Werber, X. Zhao and M. Elimelech, A facile assembled membrane with water-continuous transport
method to quantify the carboxyl group areal density in the pathways, Sci. Adv., 2019, 5, eaav9308.
This article is licensed under a Creative Commons Attribution 3.0 Unported Licence.

active layer of polyamide thin-film composite membranes, 63 X. Zhou, Y.-Y. Zhao, S.-R. Kim, M. Elimelech, S. Hu and
J. Membr. Sci., 2017, 534, 100–108. J.-H. Kim, Controlled TiO2 Growth on Reverse Osmosis and
Open Access Article. Published on 13 April 2021. Downloaded on 4/24/2023 5:23:13 PM.

56 R. Verbeke, A. Bergmaier, S. Eschbaumer, H. Mariën, Nanofiltration Membranes by Atomic Layer Deposition:


G. Dollinger and I. F. J. Vankelecom, Full elemental Mechanisms and Potential Applications, Environ. Sci. Tech-
depth-profiling with nanoscale resolution: The potential of nol., 2018, 52, 14311–14320.
Elastic Recoil Detection (ERD) in membrane science, 64 B. Liang, H. Wang, X. Shi, B. Shen, X. He, Z. A. Ghazi,
J. Membr. Sci., 2019, 572, 102–109. N. A. Khan, H. Sin, A. M. Khattak, L. Li and Z. Tang,
57 S. H. Kim, S.-Y. Kwak and T. Suzuki, Positron Annihilation Microporous membranes comprising conjugated polymers
Spectroscopic Evidence to Demonstrate the Flux- with rigid backbones enable ultrafast organic-solvent nano-
Enhancement Mechanism in Morphology-Controlled Thin- filtration, Nat. Chem., 2018, 10, 961–967.
Film-Composite (TFC) Membrane, Environ. Sci. Technol., 65 T. Wei, L. Zhang, H. Zhao, H. Ma, M. S. J. Sajib, H. Jiang and
2005, 39, 1764–1770. S. Murad, Aromatic Polyamide Reverse-Osmosis Membrane:
58 T. Fujioka, N. Oshima, R. Suzuki, W. E. Price and An Atomistic Molecular Dynamics Simulation, J. Phys.
L. D. Nghiem, Probing the internal structure of reverse Chem. B, 2016, 120, 10311–10318.
osmosis membranes by positron annihilation spectroscopy: 66 J. Muscatello, E. A. Müller, A. A. Mostofi and A. P. Sutton,
Gaining more insight into the transport of water and small Multiscale molecular simulations of the formation and
solutes, J. Membr. Sci., 2015, 486, 106–118. structure of polyamide membranes created by interfacial
59 Q. Fu, N. Verma, H. Ma, F. J. Medellin-Rodriguez, R. Li, polymerization, J. Membr. Sci., 2017, 527, 180–190.
M. Fukuto, C. M. Stafford, B. S. Hsiao and B. M. Ocko, 67 F. Foglia, S. Karan, M. Nania, Z. Jiang, A. E. Porter,
Molecular Structure of Aromatic Reverse Osmosis Polya- R. Barker, A. G. Livingston and J. T. Cabral, Neutron
mide Barrier Layers, ACS Macro Lett., 2019, 8, 352–356. Reflectivity and Performance of Polyamide Nanofilms for
60 P. S. Singh and V. K. Aswal, Compacted Nanoscale Blocks To Water Desalination, Adv. Funct. Mater., 2017, 27, 1701738.
Build Skin Layers of Reverse Osmosis and Nanofiltration 68 V. Kolev and V. Freger, Hydration, porosity and water dynamics
Membranes: A Revelation from Small-Angle Neutron in the polyamide layer of reverse osmosis membranes: A
Scattering, J. Phys. Chem. C, 2007, 111, 16219–16226. molecular dynamics study, Polymer, 2014, 55, 1420–1426.

This journal is © The Royal Society of Chemistry 2021 Chem. Soc. Rev., 2021, 50, 6290–6307 | 6307

You might also like